1: \documentclass[jcph]{apjrnl}
2:
3: \renewcommand\baselinestretch{2}
4: \usepackage{graphicx}
5: \voffset=-0.5in
6: \oddsidemargin 0pt
7: \textheight 8.8in
8: \textwidth 6.1in
9:
10: %\equationsection
11: \begin{document}
12:
13: \newcommand{\A}{{\cal A}}
14: \newcommand{\K}{{\cal K}}
15: \newcommand{\R}{{\mathbb R}}
16: \newcommand{\C}{{\mathbb C}}
17: \newcommand{\M}{{\mathbf M}}
18: \newcommand{\m}{{\mathbf m}}
19: \newcommand{\h}{{\mathbf h}}
20: \renewcommand{\H}{{\mathbf H}}
21: \newcommand{\pl}[2]{\frac{\partial#1}{\partial#2}}
22: \newcommand{\dl}[2]{\frac{\delta#1}{\delta#2}}
23: \newcommand{\p}{\partial}
24: \newcommand{\og}{\omega}
25: \newcommand{\Og}{\Omega}
26: \newcommand{\fl}[2]{\frac{#1}{#2}}
27: \newcommand{\dt}{\delta}
28: \newcommand{\td}{\tilde}
29: \newcommand{\tm}{\times}
30: \newcommand{\sm}{\setminus}
31: \newcommand{\nn}{\nonumber}
32: \newcommand{ \ap}{\alpha}
33: \newcommand{\bt}{\beta}
34: \newcommand{\ld}{\lambda}
35: \newcommand{\Gm}{\Gamma}
36: \newcommand{\gm}{\gamma}
37: \newcommand{\vp}{\varphi}
38: \newcommand{\tht}{\theta}
39: \newcommand{\ift}{\infty}
40: \newcommand{\vep}{\varepsilon}
41: \newcommand{\ep}{\epsilon}
42: \newcommand{\kp}{\kappa}
43: \newcommand{\Dt}{\Delta}
44: \newcommand{\Sg}{\Sigma}
45: \newcommand{\fa}{\forall}
46: \newcommand{\sg}{\sigma}
47: \newcommand{\ept}{\emptyset}
48: \newcommand{\btd}{\nabla}
49: \newcommand{\btu}{\bigtriangleup}
50: \newcommand{\tg}{\triangle}
51: \newcommand{\Th}{{\cal T} ^h}
52: \newcommand{\ged}{\qquad \Box}
53: \newcommand{\wt}{\widetilde}
54: \newcommand{\ol}{\overline}
55: \renewcommand{\theequation}{\arabic{section}.\arabic{equation}}
56: \newcommand{\be}{\begin{equation}}
57: \newcommand{\ee}{\end{equation}}
58: \newcommand{\ba}{\begin{array}}
59: \newcommand{\ea}{\end{array}}
60: \newcommand{\bea}{\begin{eqnarray}}
61: \newcommand{\eea}{\end{eqnarray}}
62: \newcommand{\beas}{\begin{eqnarray*}}
63: \newcommand{\eeas}{\end{eqnarray*}}
64: \newcommand{\dpm}{\displaystyle}
65:
66: \title{\bf
67: A Dynamic Atomistic-Continuum Method for the Simulation
68: of Crystalline Materials
69: \footnote{Abbreviated Title: Atomistic-Continuum Method for Crystals}}
70:
71: \author{
72: Weinan E\ref{first}\ref{second} and Zhongyi Huang\ref{first}\ref{third} \bigskip \\
73: \additem[first]{Department of Mathematics and PACM, Princeton University, Princeton, NJ 08544 \bigskip}
74: \additem[second]{School of Mathematics, Peking University, Beijing, 100866, China } \bigskip
75: \additem[third]{Department of Mathematical Sciences, Tsinghua University, Beijing, 100084, China} \bigskip
76: \\E-mail: weinan@princeton.edu, zhongyih@princeton.edu }
77:
78: \date{November 30, 2001}
79:
80: \maketitle
81:
82: \begin{abstract}
83: We present a coupled atomistic-continuum method for the modeling of
84: defects and interface dynamics of crystalline materials. The method
85: uses atomistic models such as molecular dynamics near defects and
86: interfaces, and continuum models away from defects and interfaces.
87: We propose a new class of matching conditions between the atomistic
88: and continuum regions. These conditions ensure the accurate passage
89: of large scale information between the atomistic and continuum regions
90: and at the same time minimize the reflection of phonons at the
91: atomistic-continuum interface. They can be made adaptive if we choose
92: appropriate weight functions. We present applications to dislocation
93: dynamics, friction between two-dimensional crystal surfaces and fracture
94: dynamics. We compare results of the coupled method and the detailed
95: atomistic model.
96: \end{abstract}
97:
98: \begin{keywords}
99: Atomistic-Continuum Method, Molecular Dynamics, Dislocation, Phonons, Friction,
100: Crack Propagation
101: \end{keywords}
102: \newpage
103: \tableofcontents
104: \newpage
105:
106: \section{\bf Introduction}
107: Traditionally two apparently separate approaches have been used to
108: model a continuous medium. The first is the continuum theory, in the
109: form of partial differential equations describing the conservation laws
110: and constitutive relations. This approach has been impressively successful
111: in a number of areas such as solid and fluid mechanics. It is very efficient,
112: simple and often involves very few material parameters. But it becomes
113: inaccurate for problems in which the detailed atomistic processes affect
114: the macroscopic behavior of the medium, or when the scale of the medium
115: is small enough that the continuum approximation becomes questionable.
116: Such situations are often found in studies of properties and defects of
117: micro- or nano- systems and devices. The second approach is atomistic,
118: aiming at finding the detailed behavior of each individual atom using
119: molecular dynamics or quantum mechanics. This approach can in principle
120: accurately model the underlying physical processes. But it is often
121: times prohibitively expensive.
122:
123: Recently an alternative approach has been explored that couples the
124: atomistic and continuum approaches
125: \cite{Tadmor1,Tadmor2,TimK1,TimK2,Rudd1,Rudd2,Cai}.
126: The main idea is to use atomistic modeling at places where the displacement
127: field varies on an atomic scale, and the continuum approach elsewhere.
128: The most successful and best-known implementation is the quasi-continuum
129: method \cite{Tadmor1,Tadmor2} which combines the adaptive finite element
130: procedure with an atomistic evaluation of the potential energy of the system.
131: This method has been applied to a number of examples
132: \cite{Tadmor3, Tadmor4, Tadmor5},
133: and interesting details were learned about the structure of crystal defects.
134:
135: Extension of the quasi-continuum method to dynamic problems has not been
136: straightforward \cite{Rudd1,Rudd2,Cai}. The main difficulty lies in the
137: proper matching between the atomistic and continuum regions. Since the
138: details of lattice vibrations, the phonons, which are an intrinsic part
139: of the atomistic model, cannot be represented at the continuum level,
140: conditions must be met that the phonons are not reflected at the
141: atomistic-continuum interface. Since the atomistic region is expected to be
142: a very small part of the computational domain, violation of this condition
143: quickly leads to local heating of the atomistic region and destroys the
144: simulation. In addition, the matching between the atomistic-continuum
145: interface has to be such that large scale information is accurately
146: transmitted in both directions.
147:
148: The main purpose of the present paper is to introduce a new class of matching
149: conditions between atomistic and continuum regions. These matching conditions
150: have the property that they allow accurate passage of large scale (scales
151: that are represented by the continuum model) information between the atomistic
152: and continuum regions and no reflection of phonon energy to the atomistic
153: region. These conditions can also be used in pure molecular dynamics
154: simulations as the border conditions to ensure no reflection of phonons
155: at the boundary of the simulation. As applications, we use our method to
156: study the dynamics of dislocations in the Frenkel-Kontorova model, friction
157: between crystal surfaces and crack propagation.
158:
159: \section{\bf Continuum Approximation of Atomistic Models}
160: \setcounter{equation}{0}
161: As a first step toward constructing a coupled atomistic-continuum method,
162: we discuss briefly how continuum equations are obtained from atomistic models.
163:
164: \subsection{1D Frenkel-Kontorova Model --- the Klein-Gordan Equation}
165: We first consider a simple problem, the Frenkel-Kontorova Model. This is
166: a one-dimensional chain of particles in a periodic potential,
167: coupled by springs. We will take the potential to be:
168: \begin{equation}
169: \label{FK_potential}
170: U(x)=\frac{1}{2} {\cal K} (x-a\, \mbox{int}(x/a))^2
171: \end{equation}
172: Here $a$ is the equilibrium distance between neighboring particles,
173: $\mbox{int}(x/a)$ is the integer part of $x/a$. Denote by $x_n$ the position
174: of the $n-$th particle, the dynamic equation for the particles is given by
175: \begin{equation}
176: \label{FK_force}
177: m\ddot{x}_n=k[x_{n+1}-x_n-a]+k[x_{n-1}-x_n+a]-U'(x_n) +f.
178: \end{equation}
179: where $f$ is the applied force.
180: %In the following, the spring constant $k$ is set to be 1, and so is the
181: %mass of the particles.
182:
183: One interesting aspect of the Frenkel-Kontorova model is the possibility
184: of having dislocations in the system, which corresponds to vacant
185: or doubly occupied potential wells.
186: In the absence of dislocations, the equilibrium positions of the
187: particles are given by $x_j= ja$. In general, we
188: let $x_j=a(j+u_j)$ and $\wt{f}=f/a$.
189: $u$ is then the displacement field. A dislocation corresponds to a kink
190: in $u$.
191: Far from the dislocations,
192: we can assume $|u_j - [u_j]| \ll 1$ where $[u]$ is the integer part
193: of $u$. Then we get, assuming $[u_j] = 0$,
194: \begin{equation}
195: \label{eq:FK_discrete}
196: m\ddot{u}_j=k[u_{j+1}-2u_j+u_{j-1}]-{\cal K} u_j +\wt{f}.
197: \end{equation}
198: Let $\tau=t\sqrt{m/(ka^2)}$, $\bar{\cal K}={\cal K}/(ka^{2})$, and
199: $\bar{f}=\wt{f}/(k a^2)$, we obtain:
200: \begin{equation}
201: \label{eq:FK_discrete1}
202: \pl{^2{u}_j}{\tau^2}=u_{j+1}-2u_j+u_{j-1}-\bar{\cal K} u_j +\bar{f}.
203: \end{equation}
204: Taking the limit as $a \rightarrow 0$, we obtain
205: the continuum limit equation for the displacement field $u$,
206: \be \label{FK_continuum}
207: \pl{^2 u}{\tau^2}=\pl{^2u}{x^2}-\bar{\cal K} u +\bar{f}.
208: \ee
209: This is simply the Klein-Gordan equation.
210:
211: \subsection{2D Triangular Lattice --- Isotropic Elasticity}
212:
213: Now we consider the triangular lattice model. We assume that nearest neighbor
214: atoms interact via central forces whose potential is given by
215: $\Phi(r^2)$ where $r$ is the distance between the atoms
216: (see Figure \ref{tri_lattice}). From Newton's law, we have
217: \be \label{tri_motion}
218: m \ddot{\bf r}_{\bf 0}=-\sum_{\bf j}\btd_{\bf r_{0,j}}\Phi(|{\bf r_ {0,j}}|^2),
219: \ee
220: where $m$ is the mass of the atoms, ${\bf r_j}$ is the position
221: of the ${\bf j}$-th atom (${\bf j}=(j_1,j_2)$), $\bf r_{0,j}=r_0-r_j$.
222: Let $\{\bf R_j\}$ be the equilibrium positions of the atoms. The lattice
223: constant $a$ satisfies the equilibrium condition $\Phi'(a^2)=0$. Let
224: $\{\bf u_j\}$ be the displacement vectors, ${\bf u_j = r_j - R_j}$.
225: Taylor expanding and omitting nonlinear terms in $\bf u$, we get
226: \bea
227: m\ddot{\bf u}_{\bf 0} &=& - \sum_{\bf j}\pl{}{|{\bf r_{0,j}}|^2}
228: \Phi(|{\bf r_{0,j}}|^2) \pl{|{\bf r_{0,j}}|^2}{\bf r_{0, j}} \nn \\
229: &=& -2\sum_{\bf j} \Phi'(|{\bf r_{0, j}}|^2) {\bf r_{0, j}} \nn \\
230: &=& -2\sum_{\bf j} [\Phi'(a^2)+2\Phi''(a^2){\bf R_j}\cdot({\bf u_j-u_0})]
231: [-{\bf R_j+ u_0- u_j}] \nn \\
232: &=& 4\Phi''(a^2) \sum_{\bf j} ({\bf R_j \otimes R_j}) ({\bf u_j-u_0}).
233: \label{eq:2d_tri_discrete}
234: \eea
235: Take the example of a Lennard-Jones potential:
236: \be \label{l_j_pot}
237: \Phi(r)=\ep_0\left(\fl{1}{(r/a_0)^{12}}-\fl{1}{(r/a_0)^6}\right),
238: \ee
239: then the lattice constant is equal to $a=2^{1/6}a_0$ under the assumption of
240: nearest neighbor interaction. In this case, equation (\ref{eq:2d_tri_discrete})
241: becomes
242: \bea
243: \fl{m}{\ep_0}\ddot{{\bf u}}_{\bf 0} &=&
244: \fl{18}{a^2} \left\{ \left(\ba{cc}1 & 0 \\ 0& 0 \ea \right)
245: ({\bf u}_{1,0}-2{\bf u}_{0,0}+{\bf u}_{-1,0})+
246: \left(\ba{cc} \fl{1}{4} & \fl{\sqrt{3}}{4} \\ \fl{\sqrt{3}}{4} & \fl{1}{4}\ea
247: \right) ({\bf u}_{0,1}-2{\bf u}_{0,0}+{\bf u}_{0,-1}) \right. \nn \\
248: && \left. \quad
249: +\left(\ba{cc} \fl{1}{4} & -\fl{\sqrt{3}}{4}\\ -\fl{\sqrt{3}}{4} & \fl{1}{4}\ea
250: \right) ({\bf u}_{-1,1}-2{\bf u}_{0,0}+{\bf u}_{1,-1}) \right\}.
251: \label{eq:2d_tri_discrete_lj}
252: \eea
253: Let $\tau=t\sqrt{m/\ep_0}$. The continuum limit (as $a \to 0$) of the above
254: equations is
255: \bea
256: \pl{^2 {\bf u}}{\tau^2}&=&\left(\ba{cc} 81/4 & 0 \\ 0 & 27/4 \ea\right)\pl{^2{\bf u}}{x^2}
257: +\left(\ba{cc} 27/4 & 0 \\ 0 & 81/4 \ea \right) \pl{^2{\bf u}}{y^2}
258: +\left(\ba{cc} 0 & 54/4 \\ 54/4 & 0 \ea \right) \fl{\p^2{\bf u}}{\p x\p y} \nn \\
259: &=& (\ld+\mu) \btd(\btd\cdot {\bf u}) + \mu \btu {\bf u},
260: \eea
261: where $\ld=\mu=\fl{27}{4}$. This is the equation for isotropic elasticity.
262:
263: \subsection{Slepyan Model of Fracture}
264: Here we give the one-dimensional and two-dimensional Slepyan models of
265: fracture \cite{Marder}. In the 1D case, one can view it as a model for
266: the atoms lying along a crack surface. Nearest neighbors are connected
267: by elastic springs, with spring constant $k$, and the atoms are tied to
268: the other side of the crack surface by similar springs,
269: which however snap when extended past some breaking point. The lines of
270: atoms are being pulled apart by weak springs of spring constant $k/N$.
271: These weak springs are meant schematically to represent $N$ vertical rows
272: of atoms pulling in series. Let $\{u_{j,+},u_{j,-}\}$ be the displacement
273: of atoms on the top and bottom crack surfaces respectively. The equation
274: which describes the upper line of mass points in this model is
275: \bea
276: m\ddot{u}_{j,+}=\left\{ \ba{l}
277: k(u_{j+1,+}-2u_{j,+}+u_{j-1,+}) \\
278: +\fl{k}{N}(U_N-u_{j,+}) \\
279: +k(u_{j,-}-u_{j,+})\theta(2u_f-|u_{j,-}-u_{j,+}|) \\
280: -b\dot{u}_{j,+}
281: \ea \right.
282: \label{eq:1d_slepyan_fracture_model}
283: \eea
284: Here, the first term at the right hand side is elastic coupling to neighbors,
285: the second term is the driving force by displacing edges of the strip, the
286: third term is the bonding to atoms at the opposite side of the crack surface,
287: the last term is the dissipation,
288: $\theta$ is a step function, and the
289: term containing it describes bonds which snap when their total extension
290: reaches a distance $2u_f$, where $u_f$ is a fracture distance. Assume the
291: lattice constant is $a$. In the region far away from the fracture, we have
292: \be
293: m\ddot{u}_{j,+}=k(u_{j+1,+}-2u_{j,+}+u_{j-1,+})
294: +\fl{k}{N}(U_N-u_{j,+})
295: +k(u_{j,-}-u_{j,+})
296: -b\dot{u}_{j,+}
297: \label{1d_slepyan_fracture_cont}
298: \ee
299: Dividing by $a^2$ and taking the limit as $a\to 0$, we obtain
300: \be \pl{^2u}{\tau^2}=\pl{^2u}{x^2}-\wt{b} \pl{u}{\tau}
301: \ee
302: with $\tau=t\sqrt{m/(ka^2)}$, $\wt{b}=b/(ka)$.
303:
304: Now we consider a simple 2D model (see Figure \ref{sq_lattice}). A crack moves
305: in a lattice strip composed of $2N$ rows of mass points. Assume that all the
306: atoms are located at square lattice points if there is no exterior force on
307: them. All of the bonds between lattice points are brittle-elastic, behaving as
308: perfect linear springs until the instant they snap, from which point they
309: exert no force. The displacement of each mass point is described by a single
310: spatial coordinate $u_{i,j}$, which can be interpreted as the height of mass
311: point $(i,j)$ into or out of the page. The index $i$ takes integer values,
312: while $j=1/2-N,\cdots,-1/2,1/2,\cdots,N-1/2$. The model is described by the
313: equation
314: \be \label{2d_slepyan_fracture_model}
315: m\ddot{u}_{i,j}=-b\dot{u}_{i,j}+ \sum_{\ba{c}{\rm nearest} \vspace{-4mm}\\
316: {\rm neighbors}\ (i',j')\ea} {\cal F} (u_{i',j'}-u_{i,j}),
317: \ee
318: with
319: \be \label{2d_slepyan_potential}
320: {\cal F} (r)=k r\theta(2u_f-|r|)
321: \ee
322: representing the brittle nature of the springs, $\tht$ the step function,
323: and $b$ the coefficient of a small dissipative term. The boundary condition
324: which drives the motion of the crack is
325: \be \label{bc_2d_frac}
326: u_{i,\pm(N-1/2)}=\pm U_N.
327: \ee
328: Similarly, we can get the continuum limit of (\ref{2d_slepyan_fracture_model}):
329: \be \label{2d_slepyan_fracture_cont}
330: \pl{^2 u}{\tau^2}=\btu u - \wt{b}\pl{u}{\tau}.
331: \ee
332:
333: \section{\bf Phonons}
334: \setcounter{equation}{0}
335: Among the most essential differences between the atomistic and continuum
336: behavior is the presence of phonons, the lattice vibrations, at the
337: atomistic scale. In this section we will briefly review the spectrum of
338: the phonons. Let us first consider the simplest model: 1D discrete wave
339: equation (\ref{eq:FK_discrete}) with $k=1$, ${\cal K}=0$ and $f=0$. After
340: discretization in time, we have
341: \begin{equation}
342: \label{eq:1D_discrete}
343: \frac{u^{n+1}_j-2u^n_j+u^{n-1}_j}{\Delta t^2}
344: =u^n_{j+1}-2u^n_j+u^n_{j-1}.
345: \end{equation}
346: where $u^n_j$ is the displacement of the $j$-th particle at time
347: $t=n\Delta t$.
348:
349: The phonon spectrum for (\ref{eq:1D_discrete}) is obtained by looking for
350: solutions of the type $u^n_j=e^{i(n \omega \Delta t +j \xi )}$. This gives
351: us the dispersion relation
352: \begin{equation}
353: \label{eq:1D_dispersion}
354: \frac1{\Delta t}\sin\frac{\omega\Delta t}2=\sin\fl{\xi}{2}.
355: \end{equation}
356: For the case when $\Dt t=0.01$, this dispersion relation is depicted in
357: Figure \ref{1d_dispersion}.
358: If ${\cal K} \ne 0$, we have
359: \begin{equation}
360: \label{eq:1D_dispersion_ne0}
361: \frac1{\Delta t}\sin\frac{\omega\Delta t}2=\sqrt{\sin^2\fl{\xi}{2}+{\cal K}/4}.
362: \end{equation}
363:
364: Consider now the 2D triangular lattice described by (\ref{eq:2d_tri_discrete}).
365: Let us look for the solutions of the type
366: ${\bf u}^n_{{\bf j}}=e^{i(\xi\cdot {\bf r_j} - n \omega \Delta t )} {\bf U}$
367: with ${\bf\xi}=(\xi_1,\xi_2)^T$. Substituting this expression into
368: (\ref{eq:2d_tri_discrete}), we obtain
369: \bea
370: &&\left(\fl{\sin\frac{\omega\Delta t}{2}}{\Delta t/2}\right)^2 {\bf U} \nn \\
371: &=&\fl{18}{a^2}\left[\ba{cc}
372: 4\sin^2\fl{\xi_1a}{2}+\sin^2\fl{\xi_1+\sqrt{3}\xi_2}{4}a
373: +\sin^2\fl{\xi_1-\sqrt{3}\xi_2}{4}a &
374: \sqrt{3}(\sin^2\fl{\xi_1+\sqrt{3}\xi_2}{4}a-\sin^2\fl{\xi_1-\sqrt{3}\xi_2}{4}a)\\
375: \sqrt{3}(\sin^2\fl{\xi_1+\sqrt{3}\xi_2}{4}a-\sin^2\fl{\xi_1-\sqrt{3}\xi_2}{4}a)&
376: 3(\sin^2\fl{\xi_1+\sqrt{3}\xi_2}{4}a+\sin^2\fl{\xi_1-\sqrt{3}\xi_2}{4}a) \ea
377: \right]{\bf U} \nn\\
378: &\equiv&{\bf A U}. \label{eq:2D_dispersion}
379: \eea
380: The eigenvalues of the matrix ${\bf A}$ are given by,
381: \be \label{eq:eignevalue_of_A}
382: \ld_{\pm}=\fl{36}{a^2}\left\{ \ap+\beta+\gm \pm
383: \sqrt{\ap^2+\beta^2+\gm^2-\ap\beta-\ap\gm-\beta\gm}\right\},
384: \ee
385: where
386: \[ \ap=\sin^2\fl{\xi_1}{2}a, \qquad \beta=\sin^2\fl{\xi_1+\sqrt{3}\xi_2}{4}a,
387: \qquad \gm=\sin^2\fl{\xi_1-\sqrt{3}\xi_2}{4}a.
388: \]
389: The dispersion relation now has two branches
390: \be
391: \omega_p(\xi_1,\xi_2)=\fl{2}{\Delta t}\arcsin(\fl{2}{\Delta t}\sqrt{\ld_+}),\quad
392: \omega_s(\xi_1,\xi_2)=\fl{2}{\Delta t}\arcsin(\fl{2}{\Delta t}\sqrt{\ld_-}),
393: \ee
394: where ``$p$'' and ``$s$'' stands for ``pressure'' and ``shear'' waves respectively.
395:
396: \section{\bf Optimal Local Matching Conditions}
397: \setcounter{equation}{0}
398:
399: We now come to the interface between the atomistic and continuum regions.
400: As we mentioned earlier, designing proper matching conditions at this
401: interface is a major challenge in such a coupled atomistic/continuum
402: approach.
403: The basic requirements for the matching conditions are the following:
404:
405: (1). Reflection of phonons to the atomistic region should be minimal.
406: This is particularly crucial since the atomistic regions are typically
407: very small for the purpose of computational efficiency, reflection of
408: phonon energy back to the atomistic region will trigger local heating
409: and melt the crystalline structure.
410:
411: (2). Accurate exchange of large scale information between the atomistic
412: and continuum regions.
413:
414: The first requirement is reminiscent of the absorbing boundary conditions
415: required for the computation of waves \cite{Clayton,Eng}. Indeed our work
416: draws much inspiration from that literature. There are some crucial
417: differences between the phonon problem considered here and the ones
418: studied in the literature on absorbing boundary conditions. The most
419: obvious one is the fact that the electromagnetic or acoustic waves
420: are continuum objects modeled by partial differential equations, and
421: the associated absorbing boundary conditions often use small wavenumber
422: and/or frequency approximations, whereas the phonons are intrinsically
423: discrete with substantial energy distributed at high wavenumbers.
424:
425: In the following we will give an example of a simple discrete wave
426: equation for which exact reflectionless boundary conditions can be
427: found. Such exact boundary conditions are highly nonlocal and therefore
428: not practical. But they give us guidelines on how approximate
429: boundary conditions should be constructed. We then present a method
430: that constructs optimal local matching conditions, given a predetermined
431: stencil.
432:
433: \subsection{Exact Boundary Conditions for 1D Discrete Wave Equation}
434:
435: Consider equation (\ref{eq:1D_discrete}). It is supposed to be solved for
436: all integer values of $j$. Now let us assume that we will truncate the
437: computational domain and only compute $u^n_j$ for $j\ge 0$. Assuming
438: there are no sources of waves coming from $j<0$, we still want to obtain
439: the same solution as if the computation is done for all $j$. At $j=0$,
440: we will impose a new boundary condition to make sure that the phonons
441: arriving from $j>0$ are not reflected back at $j=0$.
442:
443: At $j=0$, we replace (\ref{eq:1D_discrete}) by
444: \begin{equation}
445: \label{eq:1D_abc}
446: u^n_{0}=\sum_{k,j\ge0}a_{k,j}u^{n-k}_j, \qquad a_{0,0} = 0.
447: \end{equation}
448: We would like to determine the coefficients $\{a_{k,j}\}$. For the simple
449: problem at hand, it is possible to obtain analytical formulas of $\{a_{k,j}\}$
450: such that the imposition of (\ref{eq:1D_abc}) together with the solution of
451: (\ref{eq:1D_discrete}) for $j>0$ reproduces exactly the solution of
452: (\ref{eq:1D_discrete}) if it was solved for all integer values of $j$, i.e.
453: an exact reflectionless boundary condition can be found.
454:
455: First, let us consider the case of ${\cal K}=0$ and $f=0$. Let $\ld=\Dt t$
456: and let us look for solutions of the form:
457: \be \label{eq:1D_express}
458: u^n_j=z^n\xi^j, \qquad |\xi|\le 1. \ee
459: Substituting (\ref{eq:1D_express}) into (\ref{eq:1D_discrete}), we get
460: \be \label{eq:z_xi_relation}
461: \fl{1}{\ld^2}(z-2+\fl{1}{z}) = \xi-2+\fl{1}{\xi}.
462: \ee
463: This equation has two roots for $\xi$:
464: \be \label{eq:xi_of_z}
465: \xi_{1,2}=1+\fl{z^2-2z+1}{2\ld ^2z}\pm
466: \fl{1}{2\ld^2z}\sqrt{(z-1)^2[z^2+(4\ld^2-2)z+1]}.
467: \ee
468: Assume a boundary condition of the form
469: \be \label{eq:abc_1d_0}
470: u_0^{n+1}=2u_0^n-u_0^{n-1}+\ld^2(u_1^{n}-2u_0^n)+\sum_{k=1}^n s_k u_0^{n-k}.
471: \ee
472: Substituting (\ref{eq:1D_express}) into (\ref{eq:abc_1d_0}), we get
473: \[ z-2+\fl{1}{z}-\ld^2(\fl{1}{\xi}-2)=\sum_{k=1}^n s_k z^{-k}.\]
474: To find $s_k$, we have to find the Laurent expansion of the function
475: on the left hand side. Let
476: \be \label{eq:hz_expansion}
477: H(z)=\sqrt{(z-1)^2[z^2+(4\ld^2-2)z+1]}.
478: \ee
479: Observe that $H(z)$ satisfies
480: \[ H'(z)=2(z-1)[z^2+(3\ld^2-2)z+1-\ld^2]/H(z).\]
481: Hence
482: \be \label{eq:hz_eqn}
483: H'(z)\cdot \{(z-1)[z^2+(4\ld^2-2)z+1]\}=2(z-1)[z^2+(3\ld^2-2)z+1-\ld^2] H(z).
484: \ee
485: Solving this equation by a Laurent series:
486: $\dpm H(z)= \sum_{m\ge -2} \mu_m z^{-m}$,
487: we obtain a recursion relation $\mu_m$ for $m\ge 1$,
488: \be
489: (m+2)\mu_m=[1-2\ld^2-m(4\ld^2-3)]\mu_{m-1}
490: +[4-6\ld^2-m(3-4\ld^2)]\mu_{m-2}+(m-3)\mu_{m-3},
491: \label{eq:mu_recursion} \ee
492: and
493: \be \label{eq:mu_first}
494: \mu_{-2}=1, \quad \mu_{-1}=2\ld ^2-2, \quad \mu_0=1-2\ld^4.
495: \ee
496: Then from (\ref{eq:xi_of_z}) -- (\ref{eq:mu_first}), we have
497: \be \label{eq:1d_abc_final}
498: s_1=\ld^4, \qquad s_{k}=-\fl{\mu_{k-1}}{2}, \mbox{\ for \ } k\ge 2.
499: \ee
500: (\ref{eq:abc_1d_0}) is nonlocal and has memory effects. In order to see
501: how fast the memory decays, let us assume $\mu_k \sim m^\ap$ when $m \gg 1$,
502: substituting into (\ref{eq:mu_recursion}), and equating the coefficients of
503: term order $m^{\ap}$, we get
504: \[ 2=(4\ld^2-3)(1+\ap)+(4-6\ld^2)+2(3-4\ld^2)(1+\ap)+(2\ld^2-2)-3(1+\ap).\]
505: This gives $\ap=-2$. The decay tendency of $\mu_k$ is shown in Figure
506: \ref{fig_mu_decay}. Here $\ld=0.01$.
507:
508: If $\K \ne 0$, we can proceed as before. But (\ref{eq:xi_of_z}) changes to
509: \be \label{eq:xi_of_z_kne0}
510: \xi_{1,2}=1+\fl{\cal K}{2}+\fl{z^2-2z+1}{2\ld ^2z}\pm \fl{1}{2\ld^2z}
511: \sqrt{[z^2+({\cal K}\ld^2-2)z+1][z^2+({\cal K}\ld^2+4\ld^2-2)z+1]}.
512: \ee
513: Assuming a boundary condition of the form
514: \be \label{eq:abc_1d_kne0}
515: u_0^{n+1}=2u_0^n-u_0^{n-1}+\ld^2[u_1^{n}-(2+{\cal K})u_0^n]
516: +\sum_{k=0}^n s_k u_0^{n-k},
517: \ee
518: Substituting (\ref{eq:1D_express}) into (\ref{eq:abc_1d_kne0}), we get
519: \[ z-2+\fl{1}{z}-\ld^2(\fl{1}{\xi}-2-\K)=\sum_{k=1}^n s_k z^{-k}.\]
520: To find $s_k$, we have to find the Laurent expansion of the function on
521: the left hand side. Let $g(\K,\ld)=\K\ld^2+2\ld^2-2$ and
522: \be \label{eq:hz_expansion_kne0}
523: H(z)=\sqrt{[z^2+({\cal K}\ld^2-2)z+1][z^2+({\cal K}\ld^2+4\ld^2-2)z+1]}.
524: \ee
525: Observe that $H(z)$ satisfies
526: \[ H'(z)=\{2z^3+3\,g(\K,\ld)z^2+[g^2(\K,\ld)+2-4\ld^4]z+g(\K,\ld)\}/H(z).\]
527: Hence
528: \bea
529: &&H'(z)\cdot \{z^4+2\,g(\K,\ld)z^3+[g^2(\K,\ld)+2-4\ld^4]z^2+2\,g(\K,\ld)z+1\}
530: \nn \\ &=& H(z)\cdot\{2z^3+3\,g(\K,\ld)z^2+[g^2(\K,\ld)+2-4\ld^4]z+g(\K,\ld)\}.
531: \label{eq:hz_eqn_kne0} \eea
532: Solving this equation by a Laurent series:
533: $\dpm H(z)=\sum_{m\ge -2}\mu_m z^{-m}$,
534: we obtain a recursion relation $\mu_m$ for $m\ge 2$,
535: \bea
536: (m+2)\mu_m&=&(2m+1)[2-\ld^2(\K+2)]\mu_{m-1}
537: +(1-m)\{2-4\ld^4+[2-\ld^2(\K+2)]^2\}\mu_{m-2} \nn \\
538: && +(2-\K\ld^2-2\ld^2)(2m-5)\mu_{m-3} + (4-m)\mu_{m-4},
539: \label{eq:mu_recursion_kne0} \eea
540: and
541: \be \label{eq:mu_first_kne0}
542: \mu_{-2}=1, \quad \mu_{-1}=\K\ld^2+2\ld ^2-2, \quad \mu_0=1-2\ld^4,
543: \quad \mu_1=2\ld^4(\K\ld^2+2\ld^2-2).
544: \ee
545: Then from (\ref{eq:xi_of_z_kne0}) -- (\ref{eq:mu_first_kne0}), we have
546: \be \label{eq:1d_abc_final_kne0}
547: s_0=-\ld^2{\K},\quad s_1=\ld^4, \quad s_{k}=-\fl{\mu_{k-1}}{2},
548: \ \mbox{for } k\ge 2.
549: \ee
550: In order to see how fast the memory decays, let us assume $\mu_k \sim m^\ap$
551: when $m \gg 1$, substituting into (\ref{eq:mu_recursion_kne0}), and equating
552: the coefficients of term order $m^{\ap-1}$, we get
553: %\beas
554: %O(m^{\ap+1}): & 1=&{\cal K}({\cal K}+4) \ld^4 +1 \\
555: %O(m^{\ap}): & 2=&{\cal K}({\cal K}+4) \ld^4 (1+2\ap)+2 .
556: %\eeas
557: \[ 0=\ap\{ g(\K,\ld)(2-\ap)-2\ap[g^2(\K,\ld)+2-4\ld^2]
558: -(6+9\ap)g(\K,\ld)-8-8\ap\}. \]
559: This gives $\ap=0$ or $\ap=-2/\left\{1+\fl{\K(\K+4)}{\K+2}\ld^2\right\}$.
560:
561: These exact boundary conditions should be the same as the ones found
562: numerically in \cite{Cai}. It represents the exact Green's function
563: for (\ref{eq:1D_discrete}) which is nonlocal. However, this procedure
564: appears to be impractical for realistic models, particularly when the
565: atomistic region moves with time which is the case that interests us.
566: But such calculations can at least give us guidelines on how to proceed
567: to construct approximately reflectionless boundary conditions.
568:
569: \subsection{Optimal Local Matching Conditions for 1D Discrete Wave Equation}
570: A practical solution is to restrict (\ref{eq:1D_abc}) to a finite number of
571: terms and look for the
572: coefficients $\{a_{k,j}\}$ that minimize reflection. In order to do
573: this, let us look for solutions of
574: the type
575: \begin{equation}
576: \label{eq:1d_solu_wave}
577: u^n_j=e^{i(n\omega \Delta t+j\xi )}+R(\xi)e^{i(n\omega \Delta t -j\xi)}
578: \end{equation}
579: where $\omega$ is given by (\ref{eq:1D_dispersion}).
580: $R(\xi)$ is the reflection coefficient at wavenumber $\xi$.
581: Inserting (\ref{eq:1d_solu_wave}) into (\ref{eq:1D_abc}), we obtain
582: \begin{equation}
583: \label{eq:1d_refl_coef}
584: R(\xi)=-\frac{\sum a_{k,j}e^{i(j\xi -k\omega\Delta t)}-1}{\sum
585: a_{k,j}e^{-i(j\xi +k\omega\Delta t)}-1}
586: \end{equation}
587: The optimal coefficients $\{a_{k,j}\}$ are obtained by
588: \begin{equation}
589: \label{eq:1d_min_refl}
590: \min\int^\pi_0 W(\xi) |R(\xi)|^2d\xi
591: \end{equation}
592: subject to the constraint
593: \begin{equation}
594: \label{eq:1d_min_cons}
595: R(0)=0,R'(0)=0,\mbox{ etc.}
596: \end{equation}
597: Here $W(\xi)$ is a weight function, which is chosen to be $W(\xi) = 1$
598: in the examples below.
599:
600: Condition (\ref{eq:1d_min_cons}) guarantees that large scale information is
601: transmitted accurately, whereas (\ref{eq:1d_min_refl}) guarantees that the
602: total amount of reflection is minimized. This procedure offers a lot of
603: flexibility. For example, instead of $\int^\pi_0|R(\xi)|^2d\xi$, we can
604: minimize the total reflection over certain carefully selected interval. Another
605: possibility is to choose the weight function to be the (empirically computed)
606: energy spectrum. The coefficients $\{a_{k,j}\}$ may then change in time to
607: reflect the change of the nature of the small scales. In practice,
608: we found it preferable to use $\int^{\pi-\delta}_0|R(\xi)|^2d\xi$ with some
609: small $\delta$, instead of $\int^\pi_0|R(\xi)|^2d\xi$, in order to minimize
610: the influence of $\xi=\pi$ for which we always have $R(\pi)=1$.
611:
612: Let us look at a few examples. If in (\ref{eq:1D_abc}) we only keep the terms
613: involving $a_{1,0}$ and $a_{1,1}$, then imposing the condition $R(0)=0$ gives
614: \begin{equation}
615: \label{eq:1d_abc_first}
616: u^n_0 = (1 - \Delta t) u^{n-1}_0 +\Delta t u^{n-1}_1
617: \end{equation}
618: If instead we keep terms involving $a_{0,1}, a_{1,0}$ and $a_{1,1}$, we can
619: then impose both $R(0)=0$ and $R'(0) = 0$. This gives us
620: \begin{equation}
621: \label{eq:1d_abc_second}
622: u_0^n = u_1^{n-1} + \frac{1-\Delta t}{1+\Delta t}(u_0^{n-1} - u_1^n)
623: \end{equation}
624: Conditions of the type (\ref{eq:1d_abc_first}) and (\ref{eq:1d_abc_second}) are
625: intimately related to the absorbing boundary conditions proposed and analyzed
626: in \cite{Clayton,Eng} for the computation of waves. These conditions perform
627: well for low wavenumbers but are less satisfactory at high wavenumbers.
628:
629: To improve the performance at high wavenumbers let us consider a case that
630: include terms with $k\le2,j\le3$ and minimize $\int^{\pi-\dt}_0|R(\xi)|^2d\xi$
631: (with $\delta=0.125\pi$) subject to the condition $R(0)=0$, the optimal
632: coefficients can be easily found numerically and are given by
633: \begin{equation}
634: \label{eq:1d_abc_coef_23}
635: (a_{k,j})=\left(\begin{array}{ccc}
636: 1.95264 & -7.4207\times10^{-2} & -1.4903\times10^{-2} \\ [.15in]
637: -0.95406 & 7.4904\times10^{-2} &
638: 1.5621\times10^{-2}\end{array}\right)
639: \end{equation}
640: If instead we only include terms such that $k\le3,j\le2$, then
641: \begin{equation}
642: \label{eq:1d_abc_coef_32}
643: (a_{k,j})=\left(\begin{array}{cc}
644: 2.9524 & 1.5150\times10^{-2} \\ [.15in]
645: -2.9065 & -3.0741\times10^{-2} \\ [.15in]
646: 0.95406 & 1.5624\times10^{-2}\end{array}\right)
647: \end{equation}
648: The resulting reflection coefficients $R$ are displayed in Figure
649: \ref{reflect_coef}.
650:
651: \subsection{Optimal Local Matching Conditions for Triangular Lattice}
652: The above procedure can be easily generalized. Let us take the triangular
653: lattice as an example, and the boundary to be the $x$-axis. Given a boundary
654: condition of the form:
655: \be \label{eq:2d_abc}
656: {\bf u}^{n+1}_0 = \sum_{l\le 1} \sum_{\bf j} {\bf A}^l_{\bf j}
657: {\bf u}^{n+l}_{\bf j},
658: \ee
659: where ${\bf A}^l_{\bf j}$ are some $2\times 2$ matrices, and the summation
660: is done over a pre-selected stencil, we can find the reflection matrix
661: associated with this boundary condition. For that purpose,
662: we look for solutions of the form
663: \be \label{eq:2d_solution}
664: {\bf u}^n_{\bf j} = \sum_{\ap=I,O}\sum_{\beta=s,p} C^\ap_\beta
665: e^{i(\xi^{\ap\beta} \cdot {\bf r}_{\bf j}-\omega t)} {\bf U}^\ap_\beta,
666: \ee
667: where $\ap=I,O$ correspond to ``incoming'' and ``outgoing'' waves
668: respectively (see Figure \ref{fig_reflect}),
669: $\beta=s,p$ correspond to ``shear'' and ``pressure'' waves respectively.
670: Substituting into (\ref{eq:2D_dispersion}) we obtain a relation between
671: $(C^O_s, C^O_p)^T$ and $(C^I_s, C^I_p)^T$,
672: \be \label{eq:reflect_matrix}
673: {\bf M}_{\bf 0} \left( \ba{c}C^O_s \\ C^O_p \ea\right)
674: = {\bf M}_I\left( \ba{c}C^I_s \\ C^I_p\ea\right),
675: \ee
676: where
677: \bea \label{eq:matrix_Out}
678: {\bf M_0}&=& e^{-i\og \Dt t} [U^O_s,\ U^O_p] - \sum_{l\le 1}\sum_{\bf j}
679: {\bf A}_{\bf j}^l \left[ e^{i(\xi^{Os}\cdot{\bf r_j}-l\omega \Dt t)}U^O_s,
680: \ e^{i(\xi^{Op} \cdot{\bf r_j}-l\og \Dt t)}U^O_p\right], \\
681: \label{eq:matrix_In}
682: {\bf M}_I&=&-e^{-i\og \Dt t} [U^I_s,\ U^I_p] + \sum_{l\le 1}\sum_{\bf j}
683: {\bf A}_{\bf j}^l \left[ e^{i(\xi^{Is}\cdot{\bf r_j}-l\omega \Dt t)}U^I_s,
684: \ e^{i(\xi^{Ip} \cdot{\bf r_j}-l\og \Dt t)}U^I_p\right].
685: \eea
686: In principle, we can solve the minimization problem
687: \be \label{eq:reflect_matrix_norm}
688: \mbox{min }\int W(\xi) \|{\bf M}_I ^{-1}\cdot {\bf M_0}(\xi) \|^2 d\xi
689: \ee
690: to find optimal $\{{\bf A_j}^l\}$, where the integration is over the Brillion
691: zone. But in practice, we find it much more convenient to restrict the
692: integration over a few selected low symmetry atomic planes. In the present
693: context, it amounts to choosing special incidences where the phonons energy
694: dominates.
695:
696: First, let us consider the case of normal incidence $\tht=90^\circ$.
697: That means $\xi_1=0$. Then the matrix ${\bf A}$ in (\ref{eq:2D_dispersion})
698: becomes a diagonal matrix:
699: \beas
700: {\bf A}=\fl{18}{a^2}\left[ \ba{cc} 2\sin^2\fl{\sqrt{3}\xi_2 a}{4}& 0 \\
701: 0 & 6\sin^2\fl{\sqrt{3}\xi_2 a}{4} \ea\right].
702: \eeas
703: with two eigenvalues and eigenvectors:
704: \beas
705: \ld_1= \fl{36}{a^2}\sin^2\fl{\sqrt{3}\xi_2 a}{4}, & {\bf U}_1=(1,0)^T, \\
706: \ld_2= \fl{108}{a^2}\sin^2\fl{\sqrt{3}\xi_2 a}{4}, & {\bf U}_2=(0,1)^T.
707: \eeas
708: Then dispersion relations are
709: \bea \label{eq:thteq90_eigv1}
710: \og_s\Dt t= 2\arcsin(\fl{3\Dt t}{a}\sin\fl{\sqrt{3}\xi_2 a}{4}),
711: & {\bf U}_s=(1,0)^T, \\
712: \og_p\Dt t= 2\arcsin(\fl{3\sqrt{3}\Dt t}{a}\sin\fl{\sqrt{3}\xi_2 a}{4}),
713: & {\bf U}_p=(0,1)^T.
714: \label{eq:thteq90_eigv2}
715: \eea
716: If we take the absorb boundary condition as in (\ref{eq:2d_abc}), the matrices
717: ${\bf M_0}$ and ${\bf M_I}$ are
718: \bea \label{eq:matrix_O90}
719: {\bf M_0}&=& e^{-i\og \Dt t} I - \sum_{l\le 1}\sum_{\bf j} {\bf A}_{\bf j}^l
720: \left( \ba{cc} e^{i(\xi_2^{Os}y_{\bf j}-l\omega \Dt t)} & 0 \\
721: 0 & e^{i(\xi_2^{Op}y_{\bf j}-l\omega \Dt t)}\ea\right), \\
722: \label{eq:matrix_I90}
723: {\bf M}_I&=&-e^{-i\og \Dt t} I + \sum_{l\le 1}\sum_{\bf j} {\bf A}_{\bf j}^l
724: \left( \ba{cc} e^{i(\xi_2^{Is}y_{\bf j}-l\omega \Dt t)} & 0 \\
725: 0 & e^{i(\xi_2^{Ip}y_{\bf j}-l\omega \Dt t)}\ea\right),
726: \eea
727: where $I$ is the $2\tm 2$ identity matrix. For consistency, we should require
728: that the low wavenumber waves be transmitted accurately. Imposing
729: (\ref{eq:reflect_matrix_norm}), we get
730: \bea \label{eq:2d_constrains90_1}
731: I&=&\sum_{l\le 1} \sum_{\bf j} {\bf A} _{\bf j}^l ,\\
732: 0&=&\Dt t\ I + \sum_{l\le 1}\sum_{\bf j} {\bf A} _{\bf j}^l
733: \left( \ba{cc} \fl{2\sqrt{3}a}{9}y_{\bf j}-l\Dt t & 0 \\
734: 0 & \fl{2a}{9}y_{\bf j}-l\Dt t \ea\right)
735: \label{eq:2d_constrains90_2}\eea
736: If we minimize (\ref{eq:reflect_matrix_norm}) along normal incidence subject to
737: the constraints (\ref{eq:2d_constrains90_1}) and (\ref{eq:2d_constrains90_2}),
738: we obtain the desired matrices ${\bf A_j}^l$. For example, if we keep the terms
739: with $l=0,1$ and ${\bf j}=(0,0)$, $(-1,0)$, $(-1,1)$,
740: %(see Figure \ref{tri_lattice} and \ref{fig_reflect}),
741: the optimal coefficient matrices are
742: \beas
743: {\bf A}^0_{(0,0)}&=& \left(\ba{cc}
744: 0.947937634 & -0.423061769E-09 \\
745: -0.411523005E-09 & 0.911511476 \ea \right)\\
746: {\bf A}^0_{(-1,0)}={\bf A}^0_{(-1,1)}&=& \left(\ba{cc}
747: 0.500000011 & 0.604865049E-08 \\
748: 0.105260341E-07 & 0.499999996 \ea \right)\\
749: {\bf A}^1_{(-1,0)}={\bf A}^1_{(-1,1)}&=& \left(\ba{cc}
750: -0.473968784 & -0.603638049E-08 \\
751: -0.102331855E-07 & -0.455755718 \ea \right)
752: \eeas
753:
754: Next we consider the cases when both $\tht=60^\circ$ and $\tht=120^\circ$ are
755: taken into account. For $\tht=60^\circ$, we have $\xi_2=\sqrt{3}\xi_1$, and
756: \beas
757: {\bf A}=\fl{18}{a^2}\sin^2\fl{\xi_1 a}{2}\left[\ba{cc}9-4\sin^2\fl{\xi_1 a}{2}&
758: \sqrt{3}(3-4\sin^2\fl{\xi_1 a}{2})\\
759: \sqrt{3}(3-4\sin^2\fl{\xi_1 a}{2}) & 15-12\sin^2\fl{\xi_1 a}{2} \ea\right],
760: \eeas
761: with two eigenvalues and eigenvectors:
762: \beas \ba{rcll}
763: \ld_1&=& \fl{108}{a^2}\sin^2\fl{\xi_1 a}{2}, & {\bf U}_1=(\sqrt{3},-1)^T, \\
764: \ld_2&=& \fl{36}{a^2}\sin^2\fl{\xi_1 a}{2}(9-8\sin^2\fl{\xi_1 a}{2}),
765: & {\bf U}_2=(1,\sqrt{3})^T. \ea
766: \eeas
767: The dispersion relations are
768: \bea \label{eq:thteq60_eigv1}
769: \og_s\Dt t&=&2\arcsin\left(\fl{3\sqrt{3}\Dt t}{a}\sin\fl{\xi^s_1 a}{2}\right),
770: \qquad \qquad \qquad {\bf U}_s=(\sqrt{3},-1)^T, \\
771: \og_p\Dt t&=&2\arcsin\left(\fl{3\Dt t}{a}\sin\fl{\xi^p_1 a}{2}
772: \sqrt{9-8\sin^2\fl{\xi^p_1 a}{2}}\right), \quad {\bf U}_p=(1,\sqrt{3})^T.
773: \label{eq:thteq60_eigv2}
774: \eea
775: The consistency constraints are
776: \bea \label{eq:2d_constrains60_1}
777: I&=&\sum_{l\le 1} \sum_{{\bf j}} {\bf A}_{{\bf j}}^l ,\\
778: 0&=&\Dt t\left( \ba{cc}\sqrt{3}&1\\-1&\sqrt{3} \ea\right) +
779: \sum_{l\le 1}\sum_{\bf j} {\bf A}_{\bf j}^l \left( \ba{cc}
780: \sqrt{3}(\xi^{Is}\cdot{\bf r_j}/\omega -l\Dt t)
781: & \xi^{Ip}\cdot{\bf r_j}/\og -l\Dt t \\
782: -(\xi^{Is}\cdot{\bf r_j}/\og -l\Dt t) &
783: \sqrt{3} (\xi^{Ip}\cdot{\bf r_j}/\og -l\Dt t) \ea\right) \qquad\
784: \label{eq:2d_constrains60_2}\eea
785: for $\tht=60^\circ$, and
786: \bea \label{eq:2d_constrains120_1}
787: I&=&\sum_{l\le 1} \sum_{\bf j} {\bf A}_{\bf j}^l ,\\
788: 0&=&\Dt t\left( \ba{cc}\sqrt{3}&-1\\1&\sqrt{3} \ea\right) +
789: \sum_{l\le 1}\sum_{\bf j} {\bf A}_{\bf j}^l
790: \left( \ba{cc} \sqrt{3}(\xi^{Is}\cdot{\bf r_j}/\og -l\Dt t) &
791: -(\xi^{Ip}\cdot{\bf r_j}/\og -l\Dt t )
792: \\ \xi^{Is}\cdot{\bf r_j}/\og -l\Dt t &
793: \sqrt{3} (\xi^{Ip}\cdot{\bf r_j}/\og -l\Dt t) \ea\right) \qquad\
794: \label{eq:2d_constrains120_2}\eea
795: for $\tht=120^\circ$.
796: For example, if we keep the terms for $l=0,1$ and ${\bf j}=(0,0)$, $(-1,0)$,
797: $(-1,1)$, we have the optimal coefficient matrices
798: \beas
799: {\bf A}^0_{(0,0)}&=& \left(\ba{cc}
800: 0.929252841 & -0.861918368E-09 \\
801: 0.355336047E-09 & 0.908823412 \ea \right)\\
802: {\bf A}^0_{(-1,0)}&=& \left(\ba{cc}
803: 0.504255087 & 0.156041692E-01 \\
804: 0.161793547E-01 & 0.499201553 \ea \right)\\
805: {\bf A}^0_{(-1,1)}&=& \left(\ba{cc}
806: 0.504255087 & -0.156041692E-01 \\
807: -0.161793547E-01 & 0.499201553 \ea \right)\\
808: {\bf A}^1_{(-1,0)}&=& \left(\ba{cc}
809: -0.468881506 & 0.128308044E-01 \\
810: 0.118075167E-01 & -0.453613259 \ea \right)\\
811: {\bf A}^1_{(-1,1)}&=& \left(\ba{cc}
812: -0.468881506 & -0.128308044E-01 \\
813: -0.118075167E-01 & -0.453613259 \ea \right)
814: \eeas
815: If all three angles $\theta = 60^\circ, 90^\circ, 120^\circ$ are used with equal
816: weight, then the optimal coefficient matrices are given by:
817: \beas
818: {\bf A}^0_{(0,0)}&=& \left(\ba{cc}
819: 0.963685659E+00 & 0.522045701E-05 \\
820: 0.186532512E-05 & 0.911580620E+00
821: \ea \right)\\
822: {\bf A}^0_{(-1,0)}&=& \left(\ba{cc}
823: 0.190155146E+00 & 0.439544149E-02 \\
824: 0.132862553E-01 & 0.497292487E+00
825: \ea \right)\\
826: {\bf A}^0_{(-1,1)}&=& \left(\ba{cc}
827: 0.190158427E+00 & -0.439289996E-02 \\
828: -0.132859770E-01 & 0.497292916E+00
829: \ea \right)\\
830: {\bf A}^1_{(-1,0)}&=& \left(\ba{cc}
831: -0.171943945E+00 & 0.439598834E-02 \\
832: 0.132858254E-01 & -0.459575584E+00
833: \ea \right)\\
834: {\bf A}^1_{(-1,1)}&=& \left(\ba{cc}
835: -0.171944818E+00 & -0.439293875E-02 \\
836: -0.132856732E-01 & -0.453075576E+00
837: \ea \right)
838: \eeas
839:
840: \section{\bf Algorithms and Implementations}
841:
842: The basic framework of our coupled continuum/atomistic method is
843: that of an adaptive mesh refinement method \cite{berger}. The computational
844: domain is covered by a grid that resolves the macroscopic features of
845: problems, such as applied forces and boundary conditions. Regions near
846: atomistic defects such as dislocations, interfaces, cracks, impurities, etc
847: are detected using some error estimators. Molecular dynamics are
848: used in these regions to compute the location and momentum of each atom,
849: together with the averaged quantities at the macroscopic grid points.
850: Continuum equations are used elsewhere. At the interface between
851: the two regions, matching conditions discussed in the last section are
852: used. Specifically, we decompose the velocity and displacement
853: fields into a large scale part and a small scale part. The large
854: scale part is evolved using the values at the macroscopic grid
855: points. In the atomistic regions, these are the averaged quantities.
856: The small scale part is computed using the reflectionless
857: boundary conditions discussed above.
858:
859: One important aspect of this method is the error estimators that
860: are used to distinguish atomistic and continuum regions.
861: The senstivity of the error estimators determines the balance between
862: accuracy and efficiency. However, since there
863: has already been a lot of work done
864: on this specific problem \cite{babuska,oden,rannacher}, we will not
865: pursue this question here further. We find it adequate in our work
866: to use a refinement indicator (rather than an error estimator) which
867: is given either by an estimate of the stress, or a weighted average
868: of the wavelet coefficients.
869:
870: Further details of our method are explained through a series of
871: examples.
872:
873: \setcounter{equation}{0}
874: \subsection{Dislocation Dynamics in the Frenkel-Kontorova Model}
875: As the simplest model that encompasses most of the issues in a coupled
876: atomistic/continuum simulation, we consider the Frenkel-Kontorova model
877: \begin{equation}
878: \label{eq:1d_fk_keq0}
879: \ddot{x}_j=x_{j+1}-2x_j+x_{j-1}-U'(x_j)+f
880: \end{equation}
881: where $U$ is a periodic function with period 1, $f$ is an external forcing.
882: The continuum limit of this equation is simply the Klein-Gordan equation
883: \begin{equation} \label{eq:1d_fk_cont}
884: u_{tt}=u_{xx}-Ku+f
885: \end{equation}
886: where $K=U''(0)$. We consider the case when there is a dislocation and study
887: its dynamics under a constant applied forcing. We use $U(x)=(x-[x])^2$ where
888: $[x]$ is the integer part of $x$. In this example we take
889: (\ref{eq:1d_fk_keq0})
890: as our atomistic model, and (\ref{eq:1d_fk_cont}) as our continuum model.
891: For the coupled atomistic-continuum method, we use a standard second order
892: finite difference method for (\ref{eq:1d_fk_cont}) in the region away from
893: the dislocation, and we use (\ref{eq:1d_fk_keq0}) in the region around the
894: dislocation. However, we also place finite difference grid points in the
895: atomistic region. At these points, the values are obtained through averaging
896: the values from the atomistic model. At the interface between the atomistic
897: and continuum regions, we decompose the displacement into a large scale and
898: a small scale part. The large scale part is computed on the finite difference
899: grid, using (\ref{eq:1d_abc_coef_32}). The small scale part is computed using
900: the reflectionless boundary conditions described earlier. The interfacial
901: position between the MD and continuum regions is moved adaptively according
902: to an analysis of the wavelet coefficients or the local stress. The two
903: strategies lead to similar results. Care has to be exercised in order to
904: restrict the size of the atomistic region. For example, when wavelet
905: coefficients are used in the criteria to move the atomistic region, we found
906: it more efficient to use the intermediate levels of the wavelet coefficients
907: rather than the finest level.
908:
909: We first consider the case when a sharp transition is made between the
910: atomistic and continuum regions with a 1:16 ratio for the size of the grids.
911: Figure \ref{fig_ex_fk1} is a comparison of the displacement and velocity
912: fields computed using the full atomistic model and the coupled
913: atomistic/continuum model, with $f=0.04$. The atomistic region has 32 atoms.
914: The full atomistic simulation has $10^3$. Dislocation appears as a kink in
915: the displacement field. Notice that at the atomistic/continuum interface,
916: there is still substantial phonon energy which is then suppressed by the
917: reflectionless boundary condition. No reflection of phonons back to the
918: atomistic region is observed. In Figure \ref{dislocation_pos_04}, we compare
919: the positions of the dislocation as a function of time, computed using the
920: coupled method and the detailed molecular dynamics. Extremely good agreement
921: is observed.
922:
923: We next consider a case with $f=0.02$, which alone is too weak to move
924: the dislocation, but to the left of the dislocation, we add a sinusoidal
925: wave to the initial data. The dislocation moves as a consequence of the
926: combined effect of the force and the interaction with the wave. Yet in this
927: case the same atomistic/continuum method predicts an incorrect position for
928: the dislocation, as shown in Figure \ref{fig_ex_fk2}. The discrepancy seems to
929: grow slowly in time (see Figure \ref{dislocation_pos_02}). Improving the
930: matching conditions does not seem to lead to significant improvement.
931:
932: The difference between this case and the case shown in Figure \ref{fig_ex_fk1}
933: is that there is substantially more energy at the intermediate scales. This
934: is clearly shown in the energy spectrum that we computed for the two cases
935: but it can also be seen in Figure \ref{fig_ex_fk2} where an appreciable amount
936: of small scale waves are present in front of the dislocation.
937: Such intermediate
938: scales are suppressed in a method that uses a sharp transition between the
939: atomistic and continuum regions, unless we substantially increase the size
940: of the atomistic region. We therefore consider the next alternative
941: in which the atomistic/continuum transition is made gradually in a 1:2 or 1:4
942: ratio between neighboring grids. The right column in Figure \ref{fig_ex_fk2}
943: shows the results of such a method that uses a gradual 1:2 transition.
944: We see that the correct dislocation position is now recovered.
945:
946: \subsection{Friction between Flat and Rough Crystal Surfaces}
947: Our second example is the friction between crystal surfaces.
948: To model this process atomistically, we use standard molecular dynamics with
949: the Lennard-Jones potential \cite{Harrison,Robbins}. First, we consider the
950: case in which the two crystals are separated by a horizontal atomically
951: flat interface. The
952: atoms in the bottom crystal are assumed to be much heavier
953: (by a factor of 10) than the atoms on
954: top. To model the lack of chemical bonding between the atoms in the top
955: and bottom crystals, the interaction forces are reduced by a factor of
956: 5 between atoms in the top and bottom crystals.
957: A constant shear stress is applied near the top surface.
958: We use the periodic boundary condition
959: in the $x$-direction.
960:
961: From a physical viewpoint, one interesting issue
962: here is how dissipation takes place. Physically the kinetic energy of
963: the small scales appears as phonons which then convert into heat and
964: exit the system. A standard practice in modeling such a process is to add a
965: friction term to the molecular dynamics in order to control the temperature
966: of the system \cite{Harrison,Robbins}. In contrast, we ensure the proper
967: dissipation of phonons to the environment by imposing the reflectionless
968: boundary conditions for the phonons.
969: The results presented below are computed using the last set
970: of coefficient matrices presented at the end of Section 4.
971:
972: From Figure \ref{fig_flat_fric} we see that we indeed
973: obtain a linear relation between
974: the mean displacement of the atoms in the top crystal
975: as a function of time.
976: The temperature of the system also saturates.
977: Also plotted in Figure \ref{fig_flat_fric} is the result of the mean
978: displacement computed using the combined atomistic/continuum method. Here
979: the continuum model is the linear elastic wave
980: equation with Lame coefficients
981: computed from the Lennard-Jones potential. The agreement between the full
982: atomistic and the atomistic/continuum simulation is quite satisfactory.
983:
984: Next, we study the friction between two rough crystal surfaces. The setup
985: is the same as before, except that the initial interface between
986: the crystals
987: takes the form $y=f(x)$. The numerical results obtained are displayed
988: in Figure \ref{fig_rough_fric}.
989: In Figure \ref{friction_interface}, we plot the positions of the atoms in the top
990: and bottom crystals. We see that gaps are created in the case of
991: rough interfaces.
992:
993: In Figure \ref{vel_force}, we compare the force-velocity relations
994: for both flat and rough interfaces. Again the agreement between
995: the coupled method and the full atomistic method is quite good.
996:
997: In the present problem, we used atomistic model in a narrow strip near the
998: interface, and continuum model away from the interface. An interesting
999: question is how wide the atomistic strip has to be. Clearly for the purpose
1000: of computational efficiency, we want the atomistic strip to be as narrow as
1001: possible. On the other hand, it has to be wide enough to provide an
1002: accurate description inside the boundary
1003: layer where important atomistic processes can be relaxed. There are
1004: two important atomistic processes in the present problem. The first is the
1005: vibration of the atoms around their local equilibrium positions.
1006: The second is
1007: the process of moving from one local equilibrium to the next, i.e. sliding
1008: by one atomic distance. Clearly the second process works on longer time
1009: scale.
1010: This process has to be resolved by the atomistic layer. In Figure
1011: \ref{fig_friction_1d}, we compare the atomic positions
1012: of a column of atoms which were initially vertical, i.e. they had the same
1013: $x$-coordinates. From this picture one can also estimate the strain rate.
1014: We can clearly see that if the atomistic layer does not resolve the phonons
1015: generated by the second process, we get inaccurate results.
1016:
1017: \subsection{Crack Propagation}
1018: Our third example is the Slepyan model of fracture dynamics
1019: (\ref{eq:1d_slepyan_fracture_model}). In our coupled atomistic/continuum
1020: method, we use full atomistic simulation (\ref{eq:1d_slepyan_fracture_model})
1021: around the crack tip, and use (\ref{1d_slepyan_fracture_cont}) in the region
1022: far away from the crack tip. For the continuum equation, we use the
1023: displacement boundary
1024: condition $u_{\pm}=\pm U_N$ at the left boundary,
1025: and stress boundary condition $\frac{\partial u}{\partial x} = 0$
1026: at the right boundary.
1027: Figure \ref{fig_1d_fracture}
1028: is a comparison of the fracture surface computed using the
1029: full atomistic model
1030: and the coupled atomistic/continuum method.
1031:
1032: Next we apply our method to the 2D Mode III fracture dynamics on a
1033: square lattice (\ref{2d_slepyan_fracture_model}).
1034: Same boundary conditions as in the 1D case are used for the continuum model.
1035: For the matching conditions between the atomistic and continuum
1036: regions, we used a stencil that consists of seven points: the values
1037: of the three nearest grid points next to the boundary at the current
1038: and previous time steps, plus the value at the boundary grid point at the
1039: previous time step. The optimization is carried out using angles
1040: $\theta = 45^\circ, 90^\circ, 135^\circ$.
1041: Figure \ref{fig_2d_fracture} is a comparison
1042: of the fracture surface computed using the
1043: full atomistic model and the coupled
1044: atomistic/continuum method.
1045: Comparisons of the positions of the fracture tip as a function of
1046: time is given in Figure \ref{crack_tip}.
1047: The results are quite satisfactory.
1048: Finally in Figure \ref{shearwave}, we display the shear waves generated
1049: as a result of the crack propagation, No reflection is seen.
1050:
1051: \section{\bf Conclusion}
1052: In conclusion, we presented a new strategy for the matching conditions
1053: at the atomistic/continuum interface in multiscale modeling of crystals.
1054: The main idea is to choose the boundary condition by minimizing the reflection
1055: of phonons along a few low symmetry atomic planes, subject to some accuracy
1056: constraints at low wavenumbers. These conditions are adaptive if we choose
1057: the weight functions in (\ref{eq:1d_min_refl}) and
1058: (\ref{eq:reflect_matrix_norm}) to reflect the evolving nature of the small
1059: scales. They minimize the reflection of phonons and at the same time ensure
1060: accurate passage of large scale information. The coupled atomistic/continuum
1061: method presented here is quite robust and works well at low temperature. At
1062: finite temperature and when nonlinearity is important at large scales, a
1063: new method has to be worked out. This work is in progress.
1064:
1065:
1066: We thank Tim Kaxiras for suggesting the problem of friction
1067: between rough interface.
1068: This work is supported in part by NSF through a PECASE award and by ONR grant
1069: N00014-01-1-0674.
1070:
1071: \begin{thebibliography}{s99}
1072:
1073: \bibitem{Tadmor1} E.B. Tadmor, M. Ortiz and R. Phillips,
1074: ``Quasicontinuum analysis of defects in crystals,''
1075: \textit{Phil. Mag.}, A73 , 1529--1563 (1996).
1076:
1077: \bibitem{Tadmor2} V.B. Shenoy, R. Miller, E.B. Tadmor, D. Rodney, R.
1078: Phillips and M. Ortiz, ``An adaptive finite element approach to
1079: atomic-scale mechanics -- the quasicontinuum method,''
1080: \textit{J. Mech. Phys. Solids}, 47, 611--642 (1999).
1081:
1082: \bibitem{TimK1} F.F. Abraham, J.Q. Broughton, N. Bernstein and E.
1083: Kaxiras, ``Concurrent coupling of length scales: Methodology and
1084: application,''
1085: \textit{Phys. Rev. B}, 60 (4), 2391--2402 (1999).
1086:
1087: \bibitem{TimK2} F.F. Abraham, J.Q. Broughton, N. Bernstein and E.
1088: Kaxiras, ``Spanning the continuum to quantum length scales in a
1089: dynamic simulation of brittle fracture,''
1090: \textit{Europhys. Lett.}, 44 (6), 783--787 (1998).
1091:
1092: \bibitem{Rudd1} R.E. Rudd and J.Q. Broughton, ``Atomistic simulation of
1093: MEMS resonators through the coupling of length scales,''
1094: \textit{J. Modeling and Simulation of Microsystems}, 1 (1), 29--38 (1999).
1095:
1096: \bibitem{Rudd2} R.E. Rudd and J.Q. Broughton, ``Coarse-grained
1097: molecular dynamics and the atomic limit of finite elements,''
1098: \textit{Phys. Rev. B}, 58 (10), R5893--R5896 (1998).
1099:
1100: \bibitem{Cai} W. Cai, M. de Koning, V.V. Bulatov and S. Yip,
1101: ``Minimizing boundary reflections in coupled-domain simulations,''
1102: \textit{Phys. Rev. Lett.}, 85 (15), 3213--3216 (2000).
1103:
1104: \bibitem{Tadmor3} V.B. Shenoy, R. Miller, E.B. Tadmor, et al.
1105: ``Quasicontinuum models of interfacial structure and deformation'',
1106: \textit{Phys. Rev. Lett.}, 80 (4), 742--745 (1998).
1107:
1108: \bibitem{Tadmor4} R. Miller, E.B. Tadmor, R. Phillips R, et al.
1109: Quasicontinuum simulation of fracture at the atomic scale,
1110: \textit{Model Simul. Mater. Sc.}, 6 (5): 607--638 (1998).
1111:
1112: \bibitem{Tadmor5}G.S. Smith, E.B. Tadmor, and E. Kaxiras,
1113: ``Multiscale simulation of loading and electrical resistance in
1114: silicon nanoindentation'', \textit{Phys. Rev. Lett.} 84 (6), 1260--1263 (2000).
1115:
1116: \bibitem{EH} W. E and Z. Huang,
1117: ``Matching Conditions in Atomistic-Continuum Modeling of Materials''
1118: \textit{Phys. Rev. Lett.}, 87 (13), 135501 (2001).
1119:
1120: \bibitem{Clayton} R. Clayton and B. Engquist, ``Absorbing boundary
1121: conditions for acoustic and elastic wave equations,''
1122: \textit{Bull. Seismol. Soc. Amer.}, 67 (6), 1529--1540 (1977).
1123:
1124: \bibitem{Eng} B. Engquist and A. Majda, ``Radiation boundary
1125: conditions for acoustic and elastic wave calculations,''
1126: \textit{Comm. Pure Appl. Math.}, 32, 313--357 (1979).
1127:
1128: \bibitem{Harrison} J.A. Harrison and D.W. Brenner, \textit{Atomic-Scale
1129: simulation of tribological and related phenomena}, in ``Handbook of
1130: Micro/Nano Tribology,'' edited by B. Bhushan, CRC Press, 1995.
1131:
1132: \bibitem{Robbins} M.O. Robbins and M.H. M\"user, \textit{Computer
1133: Simulations of Friction, Lubrication and Wear}, in ``Modern Tribology
1134: Handbook,'' edited by B. Bhushan, CRC Press, 2001.
1135:
1136: \bibitem{Marder} M. Marder and S. Gross,
1137: ``Origin of crack tip instabilities'',
1138: \textit{J. Mech. Phys. Solids}, 43 (1), 1--48 (1995).
1139:
1140: \bibitem{EHuang} W. E and Z. Huang, ``Multiscale modeling of friction between
1141: crystal surfaces'', in preparation.
1142:
1143: \bibitem{berger} M. Berger and J. Oliger, ``Adaptive mesh refinement
1144: for hyperbolic partial differential equations'', \textit{J. Comput. Phys.},
1145: 53, 484 (1984).
1146:
1147: \bibitem{babuska}W.G. Szymczak and I. Babuska,
1148: ``Adaptivity and Error Estimation for the Finite-Element Method Applied to Convection Diffusion-Problems'',
1149: \textit{SIAM J. Numer. Anal.}, 21 (5), 910--954 (1984).
1150:
1151: \bibitem{oden}A. Safjan, L. Demkowicz and J.T. Oden,
1152: ``Adaptive Finite-Element Methods for Hyperbolic Systems with Application to Transient Acoustics'',
1153: \textit{Int. J. Numer. Meth. Eng.}, 32 (4), 677--707 (1991).
1154:
1155: \bibitem{rannacher}Bangerth W, Rannacher R,
1156: ``Adaptive finite element techniques for the acoustic wave equation'',
1157: \textit{J. Comput. Acoust.}, 9 (2), 575--591 (2001).
1158: \end{thebibliography}
1159:
1160: \newpage
1161: \begin{center}
1162: {\bf \large List of Figures}
1163: \end{center}
1164: \medskip
1165:
1166: \noindent{\bf FIG. \ref{tri_lattice}} %1
1167: Triagular Lattice
1168: \bigskip
1169:
1170: \noindent{\bf FIG. \ref{sq_lattice}} %2
1171: 2D Slepyan model of fracture. The white dots indicate the equilibrium locations,
1172: the black dots indicate the displaced points once stress is applied.
1173: \bigskip
1174:
1175: \noindent{\bf FIG. \ref{1d_dispersion}} %3
1176: Dispersion relation
1177: \bigskip
1178:
1179: \noindent{\bf FIG. \ref{fig_mu_decay}} %4
1180: Decay tendency of $|\mu_k|$.
1181: \bigskip
1182:
1183: \noindent{\bf FIG. \ref{reflect_coef}} %5
1184: Reflection coefficients for (\ref{eq:1d_abc_coef_23}) and
1185: (\ref{eq:1d_abc_coef_32}).
1186: \bigskip
1187:
1188: \noindent{\bf FIG. \ref{fig_reflect}} %6
1189: The `Incoming' and `Outgoing' phonons near the boundary.
1190: \bigskip
1191:
1192: \noindent{\bf FIG. \ref{fig_ex_fk1}} %7
1193: Comparison of the displacement and velocity profiles computed
1194: using the full atomistic and the atomistic/continuum models, with $f=0.04$.
1195: The top two graphs show the results in the whole computational domain.
1196: The bottom two graphs show the details near the dislocation. The solid line is the
1197: result of the atomistic/continuum method. The dash line is the result
1198: of the full atomistic method.
1199: \bigskip
1200:
1201: \noindent{\bf FIG. \ref{fig_ex_fk2}} %8
1202: Comparison of the displacement and velocity profiles computed
1203: using the full atomistic and the atomistic/continuum models, with $f=0.02$.
1204: The top two graphs show the results when the transition
1205: from the atomistic to continuum regions is sharp. The bottom two graphs
1206: show the results when the transition is gradual. Solid line is the
1207: result of the atomistic/continuum method. The dash line is the result
1208: of the full atomistic method. Only the region near the dislocation is shown.
1209: \bigskip
1210:
1211: \noindent{\bf FIG. \ref{dislocation_pos_04}} %9
1212: Comparison of the positions of the dislocation as a function of time computed
1213: using the coupled method and the detailed molecular dynamics with $f=0.04$.
1214: Dot line is the result of full MD simulation; solid line is the result with
1215: gradual transition between atomistic and continuum regions; dash line is the
1216: result with sharp transition.
1217: \bigskip
1218:
1219: \noindent{\bf FIG. \ref{dislocation_pos_02}} %10
1220: Comparison of the positions of the dislocation as a function of time computed
1221: using the coupled method and the detailed molecular dynamics with $f=0.02$.
1222: Dot line is the result of full MD simulation; solid line is the result with
1223: gradual transition between atomistic and continuum regions; dash line is the
1224: result with sharp transition.
1225: \bigskip
1226:
1227: \noindent{\bf FIG. \ref{fig_flat_fric}} %11
1228: Displacement and temperature as a function of time for the friction
1229: problem.
1230: \bigskip
1231:
1232: \noindent{\bf FIG. \ref{fig_rough_fric}} %12
1233: Displacement and temperature as a function of time for friction between
1234: rough surfaces.
1235: \bigskip
1236:
1237: \noindent{\bf FIG. \ref{friction_interface}} %13
1238: The positions of the atoms near the interfaces. The white circles are light
1239: atoms, the black ones are heavy atoms. The top graph is the initial state,
1240: the bottom graph is the late state at $t=1000$.
1241: \bigskip
1242:
1243: \noindent{\bf FIG. \ref{vel_force}} %14
1244: Comparison of the force-velocity relations for both flat and rough interfaces.
1245: The top two lines are the results for flat case,
1246: the bottom two lines are the results for rough case.
1247: The solid lines are the results for coupled atomistic/continuum method,
1248: the dash lines are the results for full MD simulation.
1249: \bigskip
1250:
1251: \noindent{\bf FIG. \ref{fig_friction_1d}} %15
1252: Comparison of the atomic positions of a column of atoms which had the same
1253: $x$-coordinates initially. Solid line is the result of full MD, the line with
1254: $\circ$ is the result of coupled method with 96 layers in the atomistic region,
1255: the line with $+$ is the result of coupled method with 16 layers in the atomistic region.
1256: In the coupled method, the ratio of atomistic and continuum grids is 1:8.
1257: \bigskip
1258:
1259: \noindent{\bf FIG. \ref{fig_1d_fracture}} %16
1260: One-dimensional fracture problem.
1261: The left graph shows the fracture surface at time $t=0$ and the
1262: right one shows the fracture surface at time $t=600$ with $b=0.01$, $N=U_N=4$.
1263: The dash line is the result of the full molecular dynamics simulation. The solid
1264: line is the result of the coupled atomistic/continuum method.
1265: The ratio of atomistic and continuum grids is 1:8.
1266: \bigskip
1267:
1268: \noindent{\bf FIG. \ref{fig_2d_fracture}} %17
1269: Two-dimensional fracture problem.
1270: The left graph shows the fracture surface at time $t=0$ and the
1271: right one shows the fracture surface at time $t=200$ with $b=0.01$, $N=512$,
1272: and $U_N=\sqrt{N}$. There are 800 atoms in each row.
1273: The dash line is the result of the full molecular dynamics simulation. The solid
1274: line is the result of the coupled atomistic/continuum method.
1275: In coupled method,
1276: we divide the whole domain into three parts. The middle part including the
1277: crack surface with $800 \tm 64$ atoms is the MD region. The top and bottom parts
1278: are continuum regions.
1279: The ratio of atomistic and continuum grids in each dimension is 1:8.
1280: \bigskip
1281:
1282: \noindent{\bf FIG. \ref{crack_tip}} %18
1283: Comparisons of the positions of the crack-tip as a function of time.
1284: The dash line is the result of the full MD simulation, the solid
1285: line is the result of the coupled method.
1286: \bigskip
1287:
1288: \noindent{\bf FIG. \ref{shearwave}} %19
1289: The shear waves.
1290: We divide the whole domain with $2048\tm 2048$ atoms into three parts.
1291: The middle part including the crack surface with $2048 \tm 128$ atoms
1292: is the MD region. The top and bottom parts are continuum regions with
1293: $128 \tm 62$ finite difference grids in each region.
1294:
1295: \noindent{\bf FIG. \ref{shearwave_elarged}} %20
1296: The enlarged picture of the MD region near the crack-tip.
1297: \bigskip
1298:
1299: \clearpage
1300: \begin{figure} %1
1301: \begin{center}
1302: \resizebox{5.5in}{!}{\includegraphics{fig_tri_lattice.ps}}
1303: \caption{}
1304: \label{tri_lattice}
1305: \end{center}
1306: \end{figure}
1307:
1308: \clearpage
1309: \begin{figure} %2
1310: \begin{center}
1311: \resizebox{5.5in}{!}{\includegraphics{sq_lattice.ps}}
1312: \caption{}
1313: \label{sq_lattice}
1314: \end{center}
1315: \end{figure}
1316:
1317: \clearpage
1318: \begin{figure} %3
1319: \begin{center}
1320: \resizebox{5.5in}{!}{\includegraphics{fig_xi_og.ps}}
1321: \caption{}
1322: \label{1d_dispersion}
1323: \end{center}
1324: \end{figure}
1325:
1326: \clearpage
1327: \begin{figure} %4
1328: \begin{center}
1329: \resizebox{5.5in}{!}{\includegraphics{fig_mu_decay.ps}}
1330: \caption{}
1331: \label{fig_mu_decay}
1332: \end{center}
1333: \end{figure}
1334:
1335: \clearpage
1336: \begin{figure} %5
1337: \begin{center}
1338: \resizebox{5.5in}{!}{\includegraphics{fig1.ps}}
1339: \caption{}
1340: \label{reflect_coef}
1341: \end{center}
1342: \end{figure}
1343:
1344: \clearpage
1345: \begin{figure} %6
1346: \begin{center}
1347: \resizebox{5.5in}{!}{\includegraphics{fig_reflect.ps}}
1348: \caption{}
1349: \label{fig_reflect}
1350: \end{center}
1351: \end{figure}
1352:
1353: \clearpage
1354: \begin{figure} %7
1355: \begin{center}
1356: \resizebox{5in}{!}{\includegraphics{fig_fk_04_1.ps}} \vspace{1cm}
1357:
1358: \resizebox{5in}{!}{\includegraphics{fig_fk_04_2.ps}}
1359: \caption{}
1360: \label{fig_ex_fk1}
1361: \end{center}
1362: \end{figure}
1363:
1364: \clearpage
1365: \begin{figure} %8
1366: \begin{center}
1367: \resizebox{5in}{!}{\includegraphics{fig_fk_02_1.ps}} \vspace{1cm}
1368:
1369: \resizebox{5in}{!}{\includegraphics{fig_fk_02_2.ps}}
1370: \caption{}
1371: \label{fig_ex_fk2}
1372: \end{center}
1373: \end{figure}
1374:
1375: \clearpage
1376: \begin{figure} %9
1377: \begin{center}
1378: \resizebox{4.5in}{!}{\includegraphics{dislocation_pos_04.ps}} \vspace{1cm}
1379:
1380: \resizebox{4.5in}{!}{\includegraphics{dislocation_pos_04_1.ps}}
1381: \caption{}
1382: \label{dislocation_pos_04}
1383: \end{center}
1384: \end{figure}
1385:
1386: \clearpage
1387: \begin{figure} %10
1388: \begin{center}
1389: \resizebox{5in}{!}{\includegraphics{dislocation_pos_02.ps}} \vspace{1cm}
1390:
1391: \resizebox{5in}{!}{\includegraphics{dislocation_pos_02_1.ps}}
1392: \caption{}
1393: \label{dislocation_pos_02}
1394: \end{center}
1395: \end{figure}
1396:
1397: \clearpage
1398: \begin{figure} %11
1399: \begin{center}
1400: \resizebox{5.5in}{!}{\includegraphics{fig4.ps}}
1401: \caption{}
1402: \label{fig_flat_fric}
1403: \end{center}
1404: \end{figure}
1405:
1406: \clearpage
1407: \begin{figure} %12
1408: \begin{center}
1409: \resizebox{5.5in}{!}{\includegraphics{fig_rough_friction.ps}}
1410: \caption{}
1411: \label{fig_rough_fric}
1412: \end{center}
1413: \end{figure}
1414:
1415: \clearpage
1416: \begin{figure} %13
1417: \begin{center}
1418: \resizebox{5in}{!}{\includegraphics{interface0.ps}} \vspace{1cm}
1419:
1420: \resizebox{5in}{!}{\includegraphics{interface.ps}}
1421: \caption{}
1422: \label{friction_interface}
1423: \end{center}
1424: \end{figure}
1425:
1426: \clearpage
1427: \begin{figure} %14
1428: \begin{center}
1429: \resizebox{5in}{!}{\includegraphics{vel_force_comp.ps}}
1430:
1431: \caption{}
1432: \label{vel_force}
1433: \end{center}
1434: \end{figure}
1435:
1436: \clearpage
1437: \begin{figure} %15
1438: \begin{center}
1439: \resizebox{5.5in}{!}{\includegraphics{fig_friction_1d_10_8.ps}}
1440: \caption{}
1441: \label{fig_friction_1d}
1442: \end{center}
1443: \end{figure}
1444:
1445: \clearpage
1446: \begin{figure} %16
1447: \begin{center}
1448: \resizebox{7.5cm}{!}{\includegraphics{fig_1d_frac0.ps}}
1449: \resizebox{7.5cm}{!}{\includegraphics{fig_1d_frac.ps}}
1450: \caption{}
1451: \label{fig_1d_fracture}
1452: \end{center}
1453: \end{figure}
1454:
1455: \clearpage
1456: \begin{figure} %17
1457: \begin{center}
1458: \resizebox{7.5cm}{!}{\includegraphics{crack0.ps}}
1459: \resizebox{7.5cm}{!}{\includegraphics{crack194.ps}}
1460: \caption{}
1461: \label{fig_2d_fracture}
1462: \end{center}
1463: \end{figure}
1464:
1465:
1466: \clearpage
1467: \begin{figure} %18
1468: \begin{center}
1469: \resizebox{5in}{!}{\includegraphics{crack_tip_comp800.ps}}
1470: \caption{}
1471: \label{crack_tip}
1472: \end{center}
1473: \end{figure}
1474:
1475: \clearpage
1476: \begin{figure} %19
1477: \begin{center}
1478:
1479: \resizebox{4in}{!}{\includegraphics{1wave480.ps}}
1480:
1481: \
1482:
1483: \resizebox{4in}{!}{\includegraphics{2wave480.ps}}
1484:
1485: \
1486:
1487: \resizebox{4in}{!}{\includegraphics{3wave480.ps}}
1488: \caption{}
1489: \label{shearwave}
1490: \end{center}
1491: \end{figure}
1492:
1493: \clearpage
1494: \begin{figure} %20
1495: \begin{center}
1496: \resizebox{4in}{!}{\includegraphics{2wave480_detail.ps}}
1497: \caption{}
1498: \label{shearwave_elarged}
1499: \end{center}
1500: \end{figure}
1501:
1502: \end{document}
1503: